Next Article in Journal
Analysis of the Mobilization of an Unsaturated Infinite Natural Slope Due to Strength Reduction
Next Article in Special Issue
Density-Driven Convection for CO2 Solubility Trapping in Saline Aquifers: Modeling and Influencing Factors
Previous Article in Journal
Influence of Using Tire-Derived Aggregate on the Structural Performance of Buried Concrete Pipe under Embankment Load
Previous Article in Special Issue
Data-Driven Modeling of Peak Rotation and Tipping-Over Stability of Rocking Shallow Foundations Using Machine Learning Algorithms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of the Effects of Freezing/Thawing Cycles on the Engineering Properties of Soilcrete

1
Department of Civil and Environmental Engineering, University of Alberta, Edmonton, AB T6G 2R3, Canada
2
Keller North America, Edmonton, AB T5S 1M3, Canada
*
Author to whom correspondence should be addressed.
Submission received: 5 October 2022 / Revised: 1 November 2022 / Accepted: 11 November 2022 / Published: 15 November 2022

Abstract

:
Soil mixing, which blends the natural soils with cementitious materials (or binders), has been used to enhance the soft ground and improve problematic soils for several decades. With developments in technique and machinery, the embedded depth of soil mixing has increased from the shallow ground to as deep as tens of meters, especially when deep soil mixing and grouting emerged in the 1970s. Extensive studies have been undertaken on the physical and mechanical properties of the mixing products (soilcrete) with regard to water content, soil type, binder type, binder content, curing age, and curing condition. However, most studies initially focused on soil mixing in temperate weather. In recent decades, soil mixing in cold regions has become common. Thus, plenty of research has been conducted on the engineering properties of soilcrete exposed to weathering conditions in cold regions, namely freezing/thawing (F/T) cycles. However, while summaries of studies on soilcrete used in temperate conditions have been undertaken by researchers, reviews of studies on soil mixing in cold regions are still rare. In order to link potential research on soil mixing with previous studies and point out the possible research directions, a review of works on soilcrete subjected to F/T cycles was composed. The present paper summarizes the testing methods adopted by various studies and the change in engineering properties of soilcrete caused by F/T cycles.

1. Introduction

The soft soil has been an obstacle for engineers and developers worldwide. Many ground improvement methods have been established since the 1920s [1,2,3,4] to deal with soft problematic soils, including shallow soil mixing, vertical drains, stone columns, and deep soil mixing (DSM). Among them, soil mixing has been used in a wide range of applications, including road pavement, subgrade, dams, shallow and deep foundations, slope, and contaminant treatment. As one of the most commonly used methods, soil mixing stands out because of the following advantages: readily available cementitious materials (or binders), relatively fast construction speed, significant improvement in strength and compressibility, and cost-effectiveness.
Soil mixing is an in situ method that blends binders with naturally incompetent soils. Generally, the binders used are lime, cement, slag, etc. Cement is the most commonly used material as it is easily made and has relatively constant components. In practice, the treated soil is abbreviated as soilcrete to represent the property changes due to the binders. Many researchers have studied the physical, chemical, and mechanical behavior of soilcrete [5,6,7]. In principle, the treatment effect is site-specific, depending on the mineralogy, deposition process, and particle size distribution of natural soils. However, most studies initially targeted the factors that influence the behavior of the soilcrete in temperate conditions, mainly in subtropical areas, because the soil mixing technique was mainly undertaken in temperate areas.
Since soil mixing has been extensively adopted in cold regions in recent decades, for example, in the applications of contaminant containment [8] and foundations for oil storage tanks [7,9], the effect of the freezing process on the soilcrete has attracted researchers’ attention, particularly when the soilcrete is anticipated to provide compressive or tensile resistance to geotechnical or structural systems. Studies have investigated the changes in the mechanical and physical behavior of soilcrete after freezing/thawing (which is often denoted as F/T) cycles in road construction, contaminant containment, foundations of superstructures and canals, and so on. So far, researchers have thoroughly documented the behavior of soilcrete in temperate conditions and have agreed on the physical and mechanical performance of soilcrete in terms of different factors, including binder content, water content, curing age, etc. However, the understanding of the engineering behavior of soilcrete subjected to F/T cycles varied according to different researchers. Thus, it becomes necessary to review the studies on soilcrete subjected to F/T cycles critically.
A thorough review of the engineering properties of soilcrete subjected to F/T cycles is conducted in the present paper. The specific objectives are as follows: (1) summarize the effects of F/T cycles on the mechanical and physical properties of soilcrete; (2) review current F/T methods and equipment; (3) present the existing understanding of the microstructure of soilcrete with and without F/T cycles; (4) propose potential future studies on soilcrete in cold regions; and (5) introduce a recent study on the effect of F/T cycles on cement-treated stiff clay used as foundation support.

2. Engineering Properties of Soilcrete in Temperate Conditions

Ordinary Portland cement is the dominant binder material in most soil mixing projects, which consists of oxides, including the main components CaO (calcium oxide) and SiO (silicon dioxide). The calcium silicate hydrate is the major product of cement hydration. Because of various chemical compositions and morphology, calcium silicate hydrate is often denoted as CSH. The soil particles or particle clusters are bonded by the CSH agglomeration or gels. As the soilcrete matrix becomes alkaline owing to the formation of calcium hydroxides (Ca(OH)2), the pozzolanic reaction occurs and generates additional CSH and calcium aluminate hydrates (CAH) [10], generating more bonds in the soilcrete. Noticeably, the CSH gels are responsible for improving the strength and stiffness of natural soils.
The binder hydration in the soil mixing usually occurs under temperate conditions, which means warm environments with an ambient temperature above 0 °C in the laboratory, on sites with a subtropical climate, or in the summer seasons of cold regions.
Many studies have been carried out on the properties of soilcrete and the performance of soil mixing projects under temperate conditions. Table 1 summarizes a few representative publications, showing the most commonly used testing methods in studies on soilcrete, such as the uniaxial compression strength (UCS) test, triaxial test, and oedometer test. Among all testing methods, the field testing has been used to identify the damage to soilcrete caused by the surroundings, and the lab testing has been used to obtain data for soil mixing design. In the beginning, core specimens obtained from the field, representing the soilcrete with in situ curing, were adopted for the lab testing [11,12]. However, the core drilling is costly, and mechanical properties of the soilcrete are necessary at the design stage. Thus, designing a proper mixing and casting procedure for the lab testing has been widely adopted.

2.1. Mechanical Properties of Soilcrete in Temperate Conditions

Products of the binder hydration enhance the mechanical properties of soilcrete under compression. Generally, the strength of soilcrete increases with curing time. Curing conditions, such as curing stress and curing temperature, also affect the strength of the soilcrete. At the early curing stage, the initial loading has been shown to remove the entrapped air in the soilcrete and greatly enhance the UCS strength of specimens [13,14,15]. Low temperature has decelerated the binder hydration, slowing down strength development [16].
Table 1 lists several representative studies that investigated the mechanical properties of soilcrete with various testing methods. The strength and stiffness are influenced by bonds formulated by the binder hydration. The UCS test is a primary method for determining the mechanical properties of soilcrete due to its reliability and speed. Based on the results of UCS testing, models considering the water/cement ratio were developed to predict the strength of soilcrete at long curing ages to facilitate the design process [17,18,19,20,21,22,23]. By comparison, Ref. [24] introduced the clay–water/cement ratio (natural clay water mass to cement mass) to determine the strength of soilcrete for fine-grained soils, which was incorporated into a prediction model and expanded the application to coarse-grained soils [25]. Additionally, Ref. [17] proposed an empirical equation for predicting the strength of soilcrete with consideration of the post-peak behavior.
Triaxial testing has also been conducted to investigate the mechanical properties of soilcrete [14,23,26,27,28,29,30]. It was concluded that the strength of soilcrete increased with confining pressure, and the performance of soilcrete resembled an overconsolidated clay, showing a strain-softening behavior. The cohesion and the friction angle have been observed to increase with cement content [27]. Many studies have implied that the preconsolidation pressure also increased with cement content [14,24,31].
The tensile strength (σt) is of great importance for soilcrete columns used for earth-retaining structures, such as slopes, dams, hydraulic barriers, runway subgrades, riverbanks, road and highway embankments, etc. Disregarding tensile forces could result in failures of soilcrete columns, as reported by [32]. Even though cement addition increases the σt of soilcrete, it does not prevent the tensile cracks from happening when subjected to external factors, such as differential settlement, wetting/drying cycles, F/T cycles, and large loads. The σt can be determined from a direct shear test, also known as the uniaxial tension test, and other testing methods. The authors of Ref. [33] determined the σt of soilcrete with the split tensile and three-point loading tests and introduced a relationship between UCS and σt.
Table 1. Key references on the performance of soilcrete under temperate conditions.
Table 1. Key references on the performance of soilcrete under temperate conditions.
ReferenceSoil TypeApplicationBindersBinder DosageTest Methods
[34]Sandy clayS/S for landfill groutCementNAUCS, permeability
[26]Sandy siltNACement and fly ashCement 2% and fly ash 4%UCS, CU, CD
[35]Soft clayColumnsCement5 to 25%UCS
[27]Silty clayColumnsCement6 to 18%CU
[10]Marine clayStabilizationCement10%, 30%, 50%XRD, SEM, MIP, PSD by laser diffractometric
[17]Soft clayStabilizationCement5 to 20%UCS, oedometer
[36]Soft clayColumnsCement8 to 33%UCS, CU, CD
[14]Marine clayStabilizationCement, lime, slag, fly ash50 to 200 kg/m3CU, CD, oedometer
[31]Marine clayStabilizationCement5 to 60%CU, CD
[37]Soft clayey soilHighway embankmentCement22.5%UCS, in situ plate loading
[38]CL (created clay and natural clay)StabilizationCement10 and 20%UCS, CIUC, CIUE, CIU Cyclic loading, oedometer
[39]MacadamHigh-speed railway subgradeCement1%, 3%, 5%Compaction, permeability, UCS
[22]CHLand reclamationCement and blast-furnace cement14% (or s/c = 7)UCS
Note: S/S—solidification/stabilization; NA—not applicable; CU—consolidated undrained; CIUC—isotropically consolidated undrained compression; CIUE—isotropically consolidated undrained extension; CIU—isotropically consolidated undrained; PSD—particle size distribution; CL—clay of low plasticity or lean clay; CH—clay of high plasticity or fat clay.

2.2. Physical Properties of Soilcrete in Temperate Conditions

A non-destructive method using elastic waves has been adopted to detect defects in soilcrete columns and obtain the elastic modulus. P and S waves propagated through the material, and the modulus of the specimen was estimated from the relationship between the travel speed of waves and the stiffness of soilcrete [40]. Moreover, the ultrasonic wave has been adopted by researchers to investigate the properties of soilcrete [41,42].
Due to the addition of cement, soilcrete has different physical properties from natural soils. It has been reported that the density of soilcrete increased with cement content [5,43]. Since compaction is often adopted to cast soilcrete specimens in many studies, dry density and optimum water content were important factors when the properties of soilcrete were analyzed. The authors of Ref. [5] observed that the maximum dry density and Atterberg limits of soilcrete increased with cement content for Iranian lean clay, but the specific gravity decreased.
The hydraulic conductivity (or permeability) of soilcrete is critical for landfill liners and deep excavation retaining structures, so determining the permeability is necessary for these studies. It has been observed that since the pores of soilcrete were filled with the binder hydration products, the permeability of soilcrete was reduced significantly [9,44,45]. The authors of Refs. [46,47] proposed a modified Moulton’s empirical equation for predicting the permeability of soilcrete by considering the effects of cement hydration and freezing. In Ref. [45], the authors showed that strength was related to permeability to a certain extent.

3. Engineering Properties of Soilcrete Subjected to F/T Cycles

Seasonal weather is typical in cold regions and detrimental to infrastructure and buildings. In light of research on the effect of seasonal weather on the behavior of concrete and natural soils, extensive studies on soilcrete subjected to F/T cycles have been undertaken. Table 2 summarizes the studies on soil mixing used in cold regions, in which the seasonal weather was simplified as F/T cycles in the lab. UCS, triaxial tests, monitoring changes in physical properties (size, water content, and mass of specimens), and permeability tests have been used to investigate the behavior of soilcrete subjected to F/T cycles. Several standards for implementing the F/T cycles have been formulated. But these standards are not suitable for soilcrete because they mainly focus on multiple cycles in a short period. Instead, a 3D F/T method with a long period in a freezer has been adopted in many studies. UCS decrease, mass loss, and size change are commonly used to indicate the detrimental effect caused by F/T cycles.

3.1. Applications of Soil Mixing in Cold Regions

As the population is increasing fast in urban areas worldwide, insufficient competent land has become an obstacle, especially in cold regions such as Japan, North America, Russia, and northern Europe. Thus, soil mixing has gradually attracted the attention of practitioners and researchers in these areas.
Figure 1 shows the typical applications of soil mixing adopted in cold regions. Soil mixing played a critical role in resisting uneven and large settlements when working as the foundations for superstructures, for example, oil storage tanks (Figure 1a) [59]. In addition, due to its increasing tensile strength stemming from the binder hydration product, soil mixing can mitigate the damage caused by the seismic load and lateral load, such as embankment (Figure 1d) [39,49,52] and retaining walls (Figure 1b). Since the permeability of soilcrete has become significantly lower than that of natural soils, soilcrete, which is often adopted in S/S, has been used to deal with contaminants (Figure 1f) [16,54,55] and reduce seepage as the dam core (Figure 1e). Soil mixing has also been used to enhance the shallow ground layer to support linear structures, such as highways [34,50], railways [39,51], and SNWTP [58].

3.2. Damage Mechanism in Soilcrete by F/T Cycles

It is generally accepted that the volume of geomaterials expands under subzero temperatures because of the phase change of water to ice. The volume increase in soils is about 9% [60]. Since soilcrete resembles natural soils, its volume change under freezing is compatible with that of soils. The freezing in soilcrete is usually unidirectional (from the ground surface to underground). As the freezing front moves downward, the ice crystals absorb the water [61], creating ice lenses in the connected pores of the soilcrete. Thus, the small pores in soilcrete are enlarged and new cracks occur. Moreover, the volume of soilcrete contracts during the subsequent thawing process. Soilcrete structures deteriorate when such an F/T reaction is repeated. For example, the permeability of soilcrete increases with the enlargement of the pores that are left after the thaw of ice crystals, as micro-fissures are enlarged and melted ice in the soilcrete matrix leaves large internal pore spaces; finer particles might move out of large pore spaces during F/T cycles [54].
Several theories for the mechanism of damages caused by F/T cycles have been summarized by [8], including the hydraulic pressure theory and the micro-ice-lens model derived from concrete studies. The hydraulic theory states that the expansion of frozen water can generate excess pore water pressure in the capillary pores. The cracks occur when the pore pressure exceeds the material’s tensile strength. In the micro-ice-lens model, on the other hand, the first F/T cycle does not damage the material significantly. Because the melting progresses from the surface to the core, the ice in the large pores remains frozen for a certain period. Thus, the ice in large pores absorbs the water from the surroundings, causing the saturation of the materials to increase notably. Hence, the damage becomes more severe with more F/T cycles.
Noticeably, the intensity of the F/T cycles plays an important role in the durability of soilcrete in cold regions, including the number of cycles, type of cycles, duration of cycles, and frost penetration depth [52].

3.3. F/T Methods and Setup

Many studies have been conducted to determine the effects of F/T cycles on the properties of concrete and natural soils. Since concrete structures exposed to air can undergo one F/T cycle per day, it is of great importance to investigate the changes in properties caused by F/T cycles [62,63,64]. In cold regions, the stability of the superstructure is reduced as a consequence of the deteriorated properties of natural soils under freezing conditions [39,58,61,65,66]. The results of Refs. [67,68] have been compiled and used to apply F/T cycles [34,50,62,64], but these standards focused on a rapid F/T process, which was not consistent with the situation for soilcrete.
Figure 2 shows two typical pieces of equipment used to conduct 3D F/T cycles in the lab for soilcrete study. In Ref. [61], the authors introduced a closed system (without a water supply) with a 3D freezing method to conduct F/T cycles in a refrigerator. The F/T processes were elongated to as long as one or two days to simulate the seasonal weather change. Recently, other similar equipments and testing methods have been adopted by studies related to soilcrete and soil subjected to F/T cycles [53,56,58]. Additionally, Ref. [69] reported that the difference between 1D and 3D freezing was negligible. Thus, a 3D freezing system with a container (refrigerator or other cells) has often been used in studies on soilcrete subjected to F/T cycles.

3.4. Physical Properties of Soilcrete Subjected to F/T Cycles

Several studies have concluded that the height of the soilcrete specimen contracted and the radius remained stable, and thus the volume decreased under freezing temperature [52,56,58]. In the warming process, changes caused by freezing may not recover fully. Because the binder hydration and F/T cycles consume water, the water content of soilcrete changes vastly during F/T cycles [50,52]. Particles at the surface of specimens were observed to be destroyed. They became loose, resulting in the cumulative mass loss (ratio of the difference between the original mass and the mass after F/T cycles to the original mass of samples) after brushing increase with F/T cycles. Mass loss has often been used to estimate the degree of damage caused by the F/T cycles in concrete [71], which has also been adopted in studies on soilcrete [34,55,57].
Figure 3 shows typical mass loss change caused by F/T cycles. Figure 3a shows that a higher water/cement ratio causes an increase in mass loss, and the accumulated mass loss after 12 cycles of F/T for cement-treated silty sands is less than 5%. Figure 3b reveals that the accumulated mass loss of lime-treated clay increased significantly with the increase in F/T cycles, which reaches about 20% for soilcrete with 3% of lime at the 12th F/T cycle. Figure 3b also indicates that the increase in lime content enhances the resistance of soilcrete to F/T cycles. Figure 3c suggests a similar result to Figure 3b that the increase in cement content improves the resistance against F/T cycles.
Permeability is important when assessing the degree of improvement of soilcrete, especially in contaminant treatment. The authors of Refs. [8,72,73] analyzed the permeability of soilcrete subjected to F/T cycles in S/S. They observed that the permeability was affected by the freezing temperature, curing time, and the number of F/T cycles. In Ref. [74], the authors suggested that the change in permeability was not caused by the pore size distribution but by the cracking and macroscale pore deformation generated when subjected to F/T cycles.
Figure 4 shows the permeability of soilcrete versus water content before and after F/T cycles. Figure 4a shows that soilcrete samples exposed to F/T cycles have higher permeability than those not exposed to F/T cycles. Figure 4b reveals that the permeability of soilcrete increases after exposure to F/T cycles, while the magnitude of the increase is smaller than those in Figure 4a. The difference in permeability shown in Figure 4 may be attributed to the soil types, the number of F/T cycles, and the procedure of F/T cycles.

3.5. Mechanical Properties of Soilcrete Subjected to F/T Cycles

The studies summarized in Table 2 show that the mechanical property of soilcrete subjected to F/T cycles is the primary concern of soil mixing practitioners. The UCS test is the first choice due to its speed and simplicity. The difference in UCS between specimens with and without F/T cycles has been used as an index to estimate the durability of soilcrete against freezing conditions [34,49,53,58]. Binders added to the natural soil could resist freezing damage to a certain extent due to the bonds generated between soil particles and clusters [49,50,56,57,76]. Generally, the strength of soilcrete decreased with F/T cycles. Researchers have observed that the most considerable change in UCS often occurred in the first or second F/T cycle [52,53,73,77]. After around ten cycles, the UCS would not change noticeably, although more cycles were applied [61,77].
Figure 5 shows the UCS of soilcrete under F/T cycles. It indicates that the UCS ratio (UCS after F/T to the original UCS) decreases significantly when F/T cycles increase from 4 to 12. The authors of Refs. [52,58] showed that the UCS of soilcrete decreases with more F/T cycles and the first cycle causes the most significant damage to the samples.
Triaxial testing has revealed that cohesion decreases with F/T cycles, while the friction angle increases in natural soils [61,66]. It is reasonable to anticipate that soilcrete would show a similar trend after F/T cycles, as the behavior of soilcrete is similar to an overconsolidated clay and studies on the triaxial behavior of soilcrete subjected to F/T cycles are scarce. In addition, detecting methods using pulse velocity [49] and resonant frequency [55] have been used to identify the damage caused by F/T cycles.
In recent decades, the use of artificial neural networks (ANNs) in geotechnical engineering studies has increased significantly, especially in modeling soil behavior, site characterization, earth retaining structures, slope stability, design of tunnels and underground openings, soil permeability, soil compaction, soil swelling, and classification of soils [78]. The authors of Ref. [79] introduced ANNs to predict the UCS of fiber-reinforced soilcrete under F/T cycles and found that ANNs performed better than the nonlinear regression method, especially the radial basis neural network approach.

3.6. Microstructure of Soilcrete Subjected to F/T Cycles

Several models of the microstructure of fine-grained soils have been proposed. Lambe’s conceptual model is widely adopted, considering clay particles as single platelets. With the advent of experimental techniques, including scanning electron microscope (SEM) and computed tomography (CT) scanning, there have been many findings on the relation between the microscale properties and the engineering behavior of fine-grained soils.
The microstructure of soilcrete is affected by binders and water, which govern the macro-behavior. For example, the porosity changes of soilcrete due to micro-cracks or cavities can be related to mechanical properties. So far, many methods have been used to investigate the microstructure of soilcrete, such as X-ray diffraction (XRD), SEM, optical microscopy, mercury intrusion porosimetry (MIP), and CT. XRD can indicate the composition of the soilcrete and has been used to estimate the degree of hydration. SEM and optical microscopy have been used to show the details of the inner structure and dispersion of hydration products inside the soilcrete specimen [80,81]. MIP can estimate the pore distribution and pore connection in the soilcrete. CT has been used to estimate the pore distribution and the porosity of soilcrete [9]. The authors of Refs. [10,82] concluded that the porosity of soilcrete was affected by the addition of cement. The microstructure of soilcrete is affected by the clay–water/cement ratio or water/binder ratio [36,83]. Thus, the strength related to microstructure could be estimated by these ratios.
Table 3 lists several key studies that investigated the microstructure of soilcrete with the new techniques. Because cement and clay interact with water in soilcrete, clay and cement particles group together when preparing soilcrete specimens. Thus, the microstructure of soilcrete was determined from the connection of the clusters and single particles in the specimens by the hydration bonds. Microstructure changes due to F/T cycles have been investigated in several studies shown in Table 3.
Soilcrete is vulnerable to F/T cycles due to the expansion of water during the freezing process. The initiation of the damage caused by F/T relies on the tensile strength of soilcrete and occurs through the weakest connections or bonds. Because of these damages, more cracks in the soilcrete specimens occur. Thus, the microstructure of soilcrete is altered by F/T processes. Many studies have been conducted on the microstructure changes in soilcrete caused by F/T cycles. Optical microscopy was used to observe the microstructure of soilcrete subjected to the F/T cycles in several studies [69,72,75], as well as other image-based methods such as SEM [56,77] and CT scanning [66]. The XRD test has been used to investigate the distribution of clay minerals in cement-treated clays. It has been confirmed by several studies that the matrix was destroyed, and the cracks were generated due to the F/T cycles. In contrast, the pore size distribution did not change much. The analysis of images could verify the strength results to a certain extent, whereas the quantitative assessment is still challenging.
Although XRD could provide the mineral amount in a powder sample of soilcrete with a specialized technique, the results may be questionable due to the small size of a single sample. Besides, the cement hydration product, i.e., CSH, is amorphous and is hard to detect with XRD. Thus, the component contents of soilcrete obtained from XRD are questionable and cross-checking with other testing methods is needed. As for the SEM and MIP methods, the sample size is relatively small compared to the cylinders for mechanical tests. Additionally, because the samples need to be dry, the sample preparation process for these tests may disturb the original microstructure, generating uncontrolled errors. With regard to SEM, a quantitative explanation of results has been lacking. Several studies introduced new methods to prepare samples with a relatively light disturbance. However, these tests are time-consuming and laborious. CT scanning is relatively new, and the results depend on soil and binder types. The major issue with the techniques mentioned above is the lack of quantitative relationships between the microstructure and the mechanical and physical properties.

3.7. Multiple Binders Used in Soilcrete Subjected to F/T Cycles

The accumulation of multiple industrial by-products, such as coal ash, silica fume, fly ash, and basanite, has become a crucial problem for practitioners. In addition, as climate change has become a major concern, plenty of research has been conducted to reduce the use of cement due to its high carbon footprint by adopting waste materials or industrial by-products. So far, many studies have been conducted on the effect of multiple binders or industrial wastes used in soil mixing.
The authors of Ref. [49] conducted UCS, California bearing ratio, ultrasonic wave, and resonant frequency tests on the cement-treated granular soils by adding silica fume, fly ash, and red mud to estimate the durability of soilcrete against F/T cycles. It was concluded that the waste materials could improve the durability of the granular soils and were suitable for road construction and earthwork applications. In Ref. [50], the authors used the UCS test to determine the optimum mixing rate of fly ash and cement for silty soils against F/T cycles with samples cast by the Standard Proctor test. The best dosage was 10% fly ash and 2% cement mixed binder. The samples after F/T cycles exhibited a crisp and brittle behavior and a higher strength which was attributed to the remarkable water loss in the curing process. The authors of Ref. [52] conducted a UCS test on soft clays stabilized with bassanite before and after exposure to F/T cycles. The strength results revealed that adding bassanite increased the strength and the durability against F/T cycles. In Ref. [70], the authors conducted a UCS test on cement-treated clays with the addition of pumice to verify the improvement of resistance against F/T cycles. It was concluded that the optimum cement content to enhance the strength was 16% and pumice could increase the resistance to F/T cycles of the cement-treated soils. Additionally, a proportion of 8% of pumice was decided as the optimum to resist F/T damage.
The authors of [53] conducted a UCS test on cement-treated clay with recycled bassanite and coal ash to investigate the resistance of soilcrete to F/T cycles. It showed that adding both additives enhanced the strength of soilcrete and the resistance to F/T cycles. In [56], the authors added cement, lime, and fly ash to the subgrade loess to improve its resistance to F/T cycles. It was concluded that the additives could prevent damage caused by F/T cycles in a certain range. There was an optimum additive content, and higher or lower dosage could hardly bear the attack by F/T cycles. The authors of Ref. [57] investigated the properties of clay blended with coal fly ash and carbide lime subjected to F/T cycles. In [57], the accumulated mass loss was used as the indicator to analyze the effects of F/T cycles, which was revealed to increase with more F/T cycles. In Ref. [76], the authors used the UCS test to investigate the effects of F/T cycles on the clay–sand mixed with cement and fly ash. It was concluded that cement-treated soil has a higher resistance to F/T cycles than fly ash-treated soil, which was attributed to the difference in the hydration and pore-filling efficiency. The increase in F/T cycles did not weaken the resistance of cement-treated soils considerably.
Recently, additives formed by the cold-bonded geopolymerization method have been adopted to stabilize the soilcrete used in grouting and deep soil mixing projects [84,85]. The authors of Ref. [84] used cold-bonded geopolymer stabilizers (consisting of metakaolin and slag) for grouting. It was found that cold-bonded stabilizers had adequate flowability and could satisfy the bearing capacity requirement for grouting. In Ref. [85], the authors deployed cold-bonded stabilizers made with limestone and bottom ash and concluded that the cold-bonded stabilizers could meet the strength requirements for grouting and deep soil mixing and provide excellent workability in projects. Thus, the cold-bonded geopolymerization is promising for grouting and deep soil mixing.
Although various additives, including coal ash, silica fume, fly ash, metakaolin, red mud, and basanite, have been introduced into the soil mixing technique to substitute cement, lower the cost, and enhance the treatment effect of cement, cement is the dominant binder in most soil mixing projects. Generally, these additives are just supplementing materials and enhance the performance of cement in the soilcrete.
Figure 6 shows several typical studies of soilcrete mixed with additives. Figure 6a concludes that adding bassanite increases the resistance of soilcrete against F/T cycles, especially at 20% usage. Meanwhile, it shows that the UCS increases with more bassanite. Figure 6b,c show that fly ash significantly affects the increase in UCS. Moreover, the combined use of fly ash and cement in natural soils significantly enhances the resistance of soilcrete to F/T cycles.

4. Research on Soilcrete with Deep Soil Mixing Method Subjected to F/T Cycles

4.1. Deep Soil Mixing Technique

In the deep soil mixing (DSM) technique, binders are blended with natural soils to strengthen the soft soil. The strengthening effect is mainly due to the reaction of binders with water. Blocks, columns, and panels are three basic patterns used in the DSM technique. Figure 7 shows the details of these patterns. Among these patterns, the cylindrical column is most common for DSM construction. Figure 8 illustrates DSM construction with a wet mixing method. The DSM columns can penetrate as deep as 70 m for offshore sites and 30 m for inland operations.
DSM was first used in the USA in 1954 [86]. Subsequently, DSM was adopted by practitioners from Japan, Europe, and America [1,2,3,4]. Lime was the first binder used by projects worldwide, and afterward, other binders were adopted, such as cement, fly ash, and slag [1,4]. Since cement hydration products have high stiffness and a strong bonding effect, cement has become predominant in DSM.
Since the 1970s, DSM has been adopted for the rapid development of offshore and onshore infrastructure, remediating the soft marine clay and other poor soils. For onshore sites, DSM has been used in the pavement and subgrade of the road system for embankments and the foundations of various structures. DSM is also an important and widely-used method in coastal projects such as docks and bridge foundations. Marine clay was the most problematic soil type in these projects, and many studies on marine clay have been conducted [10,31,87]. DSM was used in contamination treatment or S/S [88,89], land reclamation [22], and landfill liner [90]. DSM could increase strength and stiffness and reduce compressibility when dealing with soft, cohesive soils. Stiff clays may also need improvement when particular requirements are mandatory, for instance, the enormous load applied on the foundations by oil tanks [7]. The usage of DSM in cohesionless soils is not frequent. However, DSM is suitable for improving cohesionless soils in several conditions, such as mitigating the liquefaction risk and cutting off the groundwater flow. In [26], the authors mitigated the damage caused by the dilation in silt with cement. In a road subgrade project, cement strengthened the weak lateritic soils [91]. In the embankment of a high-speed railway, cement–fly ash–gravel piles were used to pass through the soft marine clay layer [92].
Several kinds of unique soils have been mixed with cement in certain areas. For example, [13,93] conducted mechanical and physical tests to investigate the behavior of cement-treated peat in Sweden.

4.2. Sample Preparation and F/T Cycles

In the present study, the specimens were made by the laboratory mixing of natural Edmonton clay and general Portland cement. Table 4 summarizes the properties of the natural Edmonton stiff clay. The compositions of natural soil and cement were examined in [9].
Most studies on soilcrete have concentrated on the low cement content (lower than 20%), but recent studies have commonly adopted high cement contents. In DSM columns installed to support oil storage tanks in eastern Edmonton, 225 kg/m3 (cement mass to the volume of natural soils and slurry) was selected to mix the soilcrete in situ. Three dosages were adopted to validate that the choice of cement content could meet the required loading conditions and study the effect of cement content on the changes in the mechanical behavior of soilcrete in the high cement range. Table 5 details the mixing plan used in the present study.
The authors of Ref. [94] suggested that tapping, rodding, dynamic compaction, static compaction, and plain mixing without compaction were five common methods to cast soilcrete specimens in the lab. The specimen preparation method developed in [94] was adopted in the present study, which was modified with the help of industrial collaborators. Figure 9 illustrates the details of the specimen preparation process. The cement and tap water were mixed with an electrical blender. The natural soil was crushed into small lumps by hand and poured into a large container. Then, the slurry was sprinkled into the lumped soil. After about 5 min, the slurry and natural soil were mixed with a dough mixer using a dough blender hook for 2 min. Then, the paste was filled into the plastic mold layer by layer. The bottom of each mold was drilled with a small hole for extruding. Before filling, the hole was sealed with waterproof tape. For each layer of filling, the mold was tapped against the hard surface of the test bench to remove the trapped air. After three layers of filling, the surface of the completed specimen was leveled. The specimens were covered with plastic film and sealed with a lid. Then they were stored in a rubber container filled with about 20 cm of tap water. The container was covered with a plastic lid to keep the moisture. The specimens were kept at room temperature (20 ± 1 °C) and left to cure for 1 to 300 days. Figure 9f–i show the cured specimens with a diameter-to-height ratio of 0.5, with diameters of 75 mm and 50 mm, respectively.
An F/T control system was designed and assembled for the present study. The equipment is located in the cold room facility with a constant room temperature of about −2 °C. Figure 10 illustrates the setup details. A steel cell circulated by spiral pipes was used as the main container, and a temperature chiller was connected to the cell by two cyclic pipes. The glycol was circulated to take away the heat from the soil. Several specimens enclosed by dry sand in the container were uniformly spread to allow the axisymmetric transition heat flows. In the present study, −20, −10, −5, and −2 °C were used to freeze the specimens. After 24 h of freezing, the chiller was set to 23 °C to warm the specimens for another 24 h, forming a complete F/T cycle.

4.3. Mechanical Testing of Soilcrete Subjected to F/T Cycles

Most studies choose a curing age of 28 and 56 days [4], representing the maturation of soilcrete specimens and the end of hydration. However, studies have shown that the development of the strength of soilcrete specimens lasted for a long time, up to two years [12,55]. It is important to learn the strength development process of soilcrete with curing time. Thus 1 to 300 days of curing was selected in the present study. After a certain curing time, the cylindrical specimens were removed from the mold through pressurized air injected into the bottom hole. Table 6 summarizes the tests conducted on the cured soilcrete specimens.
The UCS test was the most common method used by many researchers to determine the strength of the soilcrete. Before the UCS test, the specimen was capped on both ends by gypsum to obtain even surfaces for loading. The UCS test was conducted on three duplicated specimens at a strain rate of 0.5 mm/min for 75 mm specimens and about 1 mm/min for 50 mm specimens. The load and displacement were recorded until failure or 10% strain. The shear plane and failure pattern of each specimen were observed.
The CU test with the constant confining pressure was selected. Since [7] reported that soilcrete gained most of its strength in 28 days or longer, specimens for CU tests were cured for 56 days. Specimens were extruded after curing and exposed to the designed F/T cycles. They were then placed in a cylindrical cell with a designated confining limit of 4 MPa. Entrapped air in the pumps and the connection pipes were removed by water fluid. Distilled water is used in the back pump, and tap water is used in the confining pump. To ensure the saturation of soilcrete, a back pressure of 600 kPa and an initial confining pressure of 625 kPa were used and kept for 48 to 72 h. After saturation, B values were measured, all being larger than 0.9. As [38] suggested, geomaterial with a stiff microstructure was fully saturated when B was over 0.8. The permeability of the specimen was measured after saturation. The yielding pressure of soilcrete was about 1400 kPa [7]. Therefore, the saturated specimens were consolidated under different pressures varying from 100 to 1000 kPa. After consolidation, the specimen was subject to axial displacement under the undrained condition at a strain rate of 0.5% strain/hour (or 0.0125 mm/min). The specimen was compressed under an axial load until its strain reached about 15%. The pore water pressure and deviator stress were recorded every 30 s during the shearing.
The difficulties associated with the direct tension test have led to the development of several alternative indirect test methods, such as BT, the ring test, the axial fracturing test (also known as the double punch test or unconfined penetration test), bending test or flexure beam test, hydraulic fracturing test, and hoop test [95,96]. BT is common among these testing methods due to its low cost and because combining compressive stress during the test mimics the field stress. Hence, BT was chosen to determine the inter-particle or splitting strength of the soilcrete in the present study.
A thickness-to-diameter ratio of 0.5 was adopted to fabricate the specimens. A curved load strip was used in the BT, and the loading strain rate was 0.5 mm/min. The tensile strength of the BT is calculated as follows
σ t = 1.272 P m a x π t ϕ
where Pmax is the maximum applied load, t is the thickness of the specimen, and ϕ is the diameter of the specimen.
Other studies suggest that the specimen size and loading rate could greatly affect the result. Thus, it is important to ensure the specimens have a similar size and keep the loading rate steady. The tensile strength of the soilcrete was primarily related to the cohesion stemming from the clay and cement. Therefore, the tensile strength change was an indicative method to verify the microscopic changes caused by the damage to cementitious bonds.

4.4. Results of Lab Tests of Soilcrete Subjected to F/T Cycles

The soilcrete specimens were reasonably identical in terms of density, although it was difficult to prepare samples with the same mass because the mixing was not ideally uniform. The mass of C samples after curing ranged from 1207 to 1271 g. The difference in mass of most C samples was relatively marginal (<5%). The density of C samples ranged from 1.82 to 1.91 g/cm3. Similarly, the density of CS2 samples ranged from 1.57 to 1.80 g/cm3.
With an increasing F/T cycle, mass loss increased from 1.19% to 2.17% for C2 samples and from 0.88% to 9.83% for CS2 samples. As the freezing temperature decreased from −2 to −20 °C, the mass loss did not change significantly, ranging from 1 to 2%. The samples had more resistance to the mass loss for the same cement dosage but were cured for longer.
Figure 11 shows typical examples of C2 samples under different F/T cycles, indicating that mass loss is due to the cavities and cracks on the sample surface. The mass loss was progressive. As the number of F/T cycles increased, the cracks penetrated the samples, causing more clusters to burst into small debris.
Based on the UCS testing results, it was concluded that UCS deteriorated significantly because of the freezing temperature. Moreover, the decrease from −5 to −10 °C was much more significant than other temperature changes. Secondly, the decrease in UCS was progressive as the number of F/T cycles increased, exhibiting a different trend from that in [55], whose conclusion was that the first and second cycles damaged the soilcrete most, and more cycles did not continue to reduce the strength. Although the loss of strength due to the first F/T cycle was as large as 30 percent of the sample’s normal UCS without F/T cycles, the damage continued to evolve after two and more F/T cycles were applied, which is significant.
Brazilian tests were conducted with disk soilcrete specimens at a 0.5 mm/min rate. As F/T cycles damaged the bonds in the soilcrete samples, σt was heavily affected by F/T cycles. The BT results show the change in σt subjected to F/T cycles. As the freezing temperature decreased, σt decreased. When one to five F/T cycles were applied to the soilcrete, σt decreased significantly. But as 10 and 20 cycles were applied, σt became larger than in the samples without F/T cycles. The unexpected σt subjected to 10 and 20 cycles may be related to the healing effect. The authors of [97] studied the healing effect of soilcrete via the recovery in the compressive strength of soilcrete after crushing, indicating that 28 extra days of healing after crushing increased the compressive strength of soilcrete by about 27%. In [98], the authors found that the permeability of soilcrete recovered to the original value due to the healing effect of superabsorbent polymer, although the soilcrete went through aggressive environments. In these studies, the healing effect was defined as the additional cementitious products generated in soilcrete from the existing binder or external additives to cure the damaged soilcrete by filling cracks. Thus, in the aforementioned study, the long thawing period could have allowed the formation of extra hydration products, strengthening the soilcrete samples.
CT scanning images were used to estimate the overall porosity and the distribution of void ratio along soilcrete specimens [9]. Figure 12 shows that the void ratio from scanned images is close to that calculated by physical measurement. The standard deviation of the data points in Figure 12 is as small as 0.155, indicating that CT scanning is a potential method for estimating the void ratio in soilcrete. Furthermore, Figure 12 indicates that the void ratio distribution along the sample can also be obtained. The shape and volume of the cracks in the sample can be estimated by combining these photos. Thus, in the present study, CT scanning was chosen to study the microstructure of soilcrete subjected to F/T cycles.

5. Summary

According to an analysis of recent studies and the present study on DSM used in Canada, the current understanding of and potential research on soilcrete subjected to F/T cycles may be summarized as follows:
  • Soil mixing has become a common ground improvement method for practitioners and researchers worldwide, including in cold regions.
  • Mechanical and physical studies on soilcrete in temperate conditions revealed that the strength and compressibility of soilcrete have been significantly improved by adding different binders, such as fly ash, silica fume, bassanite, etc.
  • F/T cycles have been considered in many soil mixing studies where the practice was employed in cold regions. Soil mixing applications in cold regions are common, including embankments, S/S, foundations, etc.
  • F/T cycles have mostly been conducted in 3D freezers, in which the standards for concrete and compacted cement-treated clay have been adopted in several cases. Modifications of the F/T method have been conducted in many studies to meet the unique requirements for soilcrete materials used in cold regions, such as the increase in freezing time and the variation of the freezing temperature.
  • It has been generally accepted that the strength of soilcrete decreases with more F/T cycles, which the UCS could decrease by over 30% in some cases. Meanwhile, the mass loss increased with more F/T cycles, which has become an indicator of F/T destruction in soilcrete samples. Similarly, for S/S applications, the permeability of soilcrete decreased significantly after exposure to F/T cycles.
  • Although various binders have been employed in soil mixing projects, cement is the major binder used in most areas. In order to reduce the cost and deal with the increasing amount of industrial by-products or waste materials, soil mixing with new additives and industrial wastes has become popular. The studies on these binders show promising results, which can be used to substitute part of the cement, enhance the cement’s efficiency, and improve the resistance of soilcrete to F/T cycles.
  • The study on DSM used in Canada reveals that the freezing temperature significantly impacted the compressive and tensile strength of soilcrete. As temperature decreased from −2 °C to −20 °C, the magnitude of compressive strength decrease changed from 5% to about 15%. Noticeably, the damage caused by increasing F/T cycles was more important than freezing temperature.

6. Recommendations for Future Research

In summary, cement was still the most commonly used binder despite several substitutive and supplemental binders. Research on soilcrete with high cement content (over 20%) is insufficient. Consequently, UCS, triaxial tests, oedometer, and ultrasonic wave tests have been used in research to investigate the mechanical properties of soilcrete, and the permeability of soilcrete has already been determined for a few projects. So far, most studies have only focused on the compressive behavior of the soilcrete, while the tensile behavior of soilcrete has not been thoroughly understood. It has been reported that cement addition can increase the tensile strength of soilcrete, but the bonds due to cement hydration do not prevent the tensile cracks from happening when soilcrete is subjected to F/T cycles. Given that tensile strength is of great importance for soilcrete designs, particularly for soilcrete columns used for earth-retaining structures, understanding the development of soilcrete tensile strength is essential.
First, as the usage of soil mixing becomes prevalent in cold regions and thus soilcrete performance is heavily affected by F/T cycles, it is important to study the physical, hydraulic, and mechanical behavior of soilcrete subjected to F/T cycles. Moreover, the study on the performance of the soilcrete subjected to freezing temperatures becomes vital for the design and quality control/quality assessment. Although several researchers have investigated the mechanical and physical properties of soilcrete exposed to F/T cycles in contaminant treatment, road subgrade, and embankments, most studies were site-specific and concentrated on certain types of soils, and the behavior of soilcrete under severely cold weather is insufficient. Additionally, 3D freezers and standards for concrete and compacted cement-treated soil were the common methods adopted by most researchers to conduct F/T cycles. But these methods were specified for certain materials and specific freezing temperatures (depending on locality) and were not universal, resulting in difficulty in comparing the results from different studies. The need to formulate a universal method for employing F/T cycles is becoming urgent.
Secondly, almost all studies concentrated on the effect of the number of F/T cycles on the properties of soilcrete, while the freezing temperature was not taken into consideration. Remarkably, temperature plays an important role in determining the speed of the freezing process, which is assumed to be the major direct cause of the F/T damage. Thus, studying the effect of freezing temperature becomes necessary.
Thirdly, in almost all studies on soilcrete subjected to F/T cycles, the major reason for the damage was assumed to be the volume change due to water transforming to ice. However, the variation in volume change from different minerals may also contribute to the cracks and cavities from F/T cycles. In order to thoroughly understand the mechanism of F/T damage, the study of dry soilcrete samples subjected to F/T cycles should attract researchers’ attention.
Moreover, as addressed in past studies, the pozzolanic reaction between clay minerals and the binders occurs in soilcrete at a late stage of curing. It can strengthen the soilcrete to some extent based on the amount of the binder hydration product. However, several types of soils contain minerals other than clay minerals, so the potential reactions between these minerals and the binder are unclear.
In addition, many studies have investigated the microstructure changes in soilcrete caused by the binder hydration and F/T cycles. However, there is no overall pattern for all types of soils treated with the same binder, and the findings from different studies vary depending on the soil type, binder type, and casting methods. Furthermore, the studies on microstructure are mostly qualitative. Thus, it is necessary to develop a method to quantify the microstructure of soilcrete and unify the damage patterns in soilcrete caused by F/T cycles.
Finally, since conduction dominates the heat transfer in soilcrete and natural soils, thermal conductivity becomes the most important factor determining the temperature distribution in soilcrete samples. However, information on the thermal properties of soilcrete is scarce, especially under freezing temperatures.

Author Contributions

Conceptualization: C.L., C.B. and L.D.; writing—original draft preparation: C.L.; writing—review and editing: C.L., C.B. and L.D.; supervision: L.D.; funding acquisition: L.D. All authors have read and agreed to the published version of the manuscript.

Funding

The research is funded by the Natural Sciences and Engineering Research Council of Canada (NSERC) under the Collaborative R&D program (CRDPJ 493088-2015).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Porbaha, A. State of the Art in Deep Mixing Technology Part I. Basic Concepts and Overview. Ground Improv. 1998, 2, 81–92. [Google Scholar] [CrossRef]
  2. Terashi, M. The State of Practice in Deep Mixing Methods. In Grouting and Ground Treatment; ASCE: Reston, VA, USA, 2003; pp. 25–49. [Google Scholar]
  3. Larsson, S. State of Practice Report—Execution, Monitoring and Quality Control. In Proceedings of the International Conference on Deep Mixing Best Practice and Recent Advances–Deep Mixing, Stockholm, Sweden, 23–25 May 2005; pp. 732–785. [Google Scholar]
  4. Han, J. Principles and Practice of Ground Improvement; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015. [Google Scholar]
  5. Pakbaz, M.S.; Alipour, R. Influence of Cement Addition on the Geotechnical Properties of an Iranian Clay. Appl. Clay Sci. 2012, 67–68, 1–4. [Google Scholar] [CrossRef]
  6. Eskisar, T. Influence of Cement Treatment on Unconfined Compressive Strength and Compressibility of Lean Clay with Medium Plasticity. Arab. J. Sci. Eng. 2015, 40, 763–772. [Google Scholar] [CrossRef]
  7. Luis, A.; Deng, L. Development of Mechanical Properties of Edmonton Stiff Clay Treated with Cement and Fly Ash. Int. J. Geotech. Eng. 2018, 14, 329–339. [Google Scholar] [CrossRef]
  8. Jamshidi, R.J. Evaluation of Cement-Treated Soils Subjected to Cycles of Freezing and Thawing. Ph.D. Thesis, Dalhousie University, Halifax, NS, Canada, 2014. [Google Scholar]
  9. Luis, A.; Deng, L.; Shao, L.; Li, H. Triaxial Behaviour and Image Analysis of Edmonton Clay Treated with Cement and Fly Ash. Constr. Build. Mater. 2019, 197, 208–219. [Google Scholar] [CrossRef]
  10. Chew, S.H.; Kamruzzaman, A.H.M.; Lee, F.H. Physicochemical and Engineering Behavior of Cement Treated Clays. J. Geotech. Geoenviron. Eng. 2004, 130, 696–706. [Google Scholar] [CrossRef]
  11. Fang, Y.S.; Chung, Y.T.; Chen, T.J. Properties of Soil-Cement Stabilised with Deep Mixing Method. Ground Improv. 2001, 5, 69–74. [Google Scholar] [CrossRef]
  12. Ikegami, M.; Ichiba, T.; Ohishi, K.; Terashi, M. Long-Term Properties of Cement Treated Soil 20 Years after Construction. In Proceedings of the 16th International Conference on Soil Mechanics and Geotechnical Engineering, Osaka, Japan, 12–16 September 2005; pp. 1199–1202. [Google Scholar]
  13. Ahnberg, H.; Bengtsson, P.E.; Holm, G. Effect of Initial Loading on the Strength of Stabilised Peat. Ground Improv. 2001, 5, 35–40. [Google Scholar] [CrossRef]
  14. Ahnberg, H. Consolidation Stress Effects on the Strength of Stabilised Swedish Soils. Ground Improv. 2006, 10, 1–13. [Google Scholar] [CrossRef]
  15. Liu, C.; Starcher, R.D. Effects of Curing Conditions on Unconfined Compressive Strength of Cement-and Cement-Fiber-Improved Soft Soils. J. Mater. Eng. 2013, 25, 1134–1141. [Google Scholar] [CrossRef]
  16. Rothhämel, M.; Laue, J. Influence of Cold Curing Temperature and Freeze–Thaw on the UCS of Stabilised Silty Sand. Proc. Inst. Civil Eng. Ground Improv. 2020, 174, 232–239. [Google Scholar] [CrossRef]
  17. Lorenzo, G.A.; Bergado, D.T. Fundamental Characteristics of Cement-Admixed Clay—New Approach. J. Geotech. Geoenviron. Eng. 2004, 130, 1042–1050. [Google Scholar] [CrossRef]
  18. Horpibulsuk, S.; Rachan, R.; Suddeepong, A. Assessment of Strength Development in Blended Cement Admixed Bangkok Clay. Constr. Build. Mater. 2011, 25, 1521–1531. [Google Scholar] [CrossRef]
  19. Consoli, N.C.; Rosa, D.A.; Cruz, R.C.; Rosa, A.D. Water Content, Porosity and Cement Content as Parameters Controlling Strength of Artificially Cemented Silty Soil. Eng. Geol. 2011, 122, 328–333. [Google Scholar] [CrossRef]
  20. He, L. Spatial Variability in Deep Soil Mixing—Centrifuge Model and Field Data Study. Ph.D. Thesis, National University of Singapore, Singapore, 2018. [Google Scholar]
  21. Yao, K.; Pan, Y.; Jia, L.; Yi, J.T.; Hu, J.; Wu, C. Strength Evaluation of Marine Clay Stabilized by Cementitious Binder. Mar. Georesour. Geotechnol. 2019, 38, 730–743. [Google Scholar] [CrossRef]
  22. Bi, J.; Chian, S.C. Modelling of Three-Phase Strength Development of Ordinary Portland Cement- and Portland Blast-Furnace Cement-Stabilised Clay. Geotechnique 2020, 70, 80–89. [Google Scholar] [CrossRef]
  23. Ho, T.-O.; Chen, W.-B.; Yin, J.-H.; Wu, P.-C.; Tsang, D.C.W. Stress-Strain Behaviour of Cement-Stabilized Hong Kong Marine Deposits. Constr. Build. Mater. 2021, 274, 122103. [Google Scholar] [CrossRef]
  24. Miura, N.; Horpibulsuk, S.; Nagaraj, T.S. Engineering Behavior of Cement Stabilized Clay at High Water Content. Soils Found. 2001, 41, 33–45. [Google Scholar] [CrossRef] [Green Version]
  25. Horpibulsuk, S.; Katkan, W.; Sirilerdwattana, W.; Rachan, R. Strength Development in Cement Stabilized Low Plasticity and Coarse Grained Soils- Laboratory and Field Study. Soils Found. 2006, 46, 351–366. [Google Scholar] [CrossRef] [Green Version]
  26. Lo, S.R.; Wardani, S.P. Strength and Dilatancy of a Silt Stabilized by a Cement and Fly Ash Mixture. Can. Geotech. J. 2002, 39, 77–89. [Google Scholar] [CrossRef]
  27. Horpibulsuk, S.; Miura, N.; Bergado, D.T. Undrained Shear Behavior of Cement Admixed Clay at High Water Content. J. Geotech. Geoenviron. Eng. 2004, 130, 1096–1105. [Google Scholar] [CrossRef]
  28. Ahnber, H. Effects of Back Pressure and Strain Rate Used in Triaxial Testing of Stabilized Organic Soils and Clays. Geotech. Test. J. 2004, 27, 250–259. [Google Scholar]
  29. Kasama, K.; Zen, K.; Iwataki, K. Undrained Shear Strength of Cement-Treated Soils. Soils Found. 2006, 46, 221–232. [Google Scholar] [CrossRef] [Green Version]
  30. Shen, S.; Han, J.; Du, Y. Deep Mixing Induced Property Changes in Surrounding Sensitive Marine Clays. J. Geotech. Geoenviron. Eng. 2008, 134, 845–854. [Google Scholar] [CrossRef]
  31. Kamruzzaman, A.H.M.; Chew, S.H.; Lee, F.H. Structuration and Destructuration Behavior of Cement-Treated Singapore Marine Clay. J. Geotech. Geoenviron. Eng. 2009, 135, 573–589. [Google Scholar] [CrossRef]
  32. Lee, S.A. Characterization and Modeling of Cement—Treated Soil Column Used as Cantilever Earth Retaining Structure. Ph.D. Thesis, National University of Singapore, Singapore, 2014. [Google Scholar]
  33. Ng, K.S.; Chew, Y.M.; Nordin, S. Fracture Behaviour of Cement Treated Sandy Clay. In Proceedings of the MATEC Web Conference-3rd International Conference on Building Materials and Materials Engineering, Lisbon, Portugal, 25–27 September 2019; Volume 303, p. 01001. [Google Scholar]
  34. Shihata, S.A.; Baghdadi, Z.A. Simplified Method to Assess Freeze-Thaw Durability of Soil Cement. J. Mater. Eng. 2001, 13, 243–247. [Google Scholar] [CrossRef]
  35. Horpibulsuk, S.; Rachan, R.; Suddeepong, A. Assessment of Strength Development in Cement-Admixed High Water Content Clays with Abrams’ Law as a Basis. Geotechnique 2003, 53, 439–444. [Google Scholar] [CrossRef]
  36. Horpibulsuk, S.; Miura, N.; Nagaraj, T.S. Clay–Water/Cement Ratio Identity for Cement Admixed Soft Clays. J. Geotech. Geoenviron. Eng. 2005, 131, 187–192. [Google Scholar] [CrossRef]
  37. Liu, S.-Y.; Du, Y.-J.; Yi, Y.-L.; Puppala, A.J. Field Investigations on Performance of T-Shaped Deep Mixed Soil Cement Column–Supported Embankments over Soft Ground. J. Geotech. Geoenviron. Eng. 2012, 138, 718–727. [Google Scholar] [CrossRef]
  38. Quiroga, A.J.; Thompson, Z.M.; Muraleetharan, K.K.; Miller, G.A.; Cerato, A.B. Stress–Strain Behavior of Cement-Improved Clays: Testing and Modeling. Acta Geotech. 2017, 12, 1003–1020. [Google Scholar] [CrossRef]
  39. Wang, T.; Wang, H.; Song, H.; Yue, Z.; Guo, Z. Effects of Cement Content and Grain-Size Composition on Engineering Properties of High-Speed-Railway Macadam Subgrade. Cold Reg. Sci. Technol. 2018, 145, 21–31. [Google Scholar] [CrossRef]
  40. Kokusho, T. Innovative Earthquake Soil Dynamics; CRC Press, Taylor & Francis Group, Inc.: Boca Raton, FL, USA, 2017. [Google Scholar]
  41. Panzera, T.H.; Christoforo, A.L.; Cota, F.P.; Borges, P.H.R.; Bowen, C.R. Ultrasonic Pulse Velocity Evaluation of Cementitious Materials. In Advances in Composite Materials—Analysis of Natural and Man-Made Materials; Tesinova, P., Ed.; IntechOpen: London, UK, 2011; pp. 411–436. [Google Scholar]
  42. Åhnberg, H.; Holmén, M. Assessment of Stabilised Soil Strength with Geophysical Methods. Proc. Inst. Civil Eng. Ground Improv. 2011, 164, 109–116. [Google Scholar] [CrossRef]
  43. Al-Jabban, W.J. Soil Modification by Adding Small Amounts of Soil Stabilizers Impact of Portland Cement and the Industrial By-Product Petrit T. Master’s Thesis, Luleå University of Technology, Luleå, Sweden, 2017. [Google Scholar]
  44. Ahnberg, H. Measured Permeabilities in Stabilized Swedish Soils. In Grouting and Ground Treatment; ASCE: Reston, VA, USA, 2003; pp. 622–633. [Google Scholar]
  45. Bellezza, I.; Frataloccchi, E. Effectiveness of Cement on Hydraulic Conductivity of Compacted Soil–Cement Mixtures. Ground Improv. 2006, 10, 77–90. [Google Scholar] [CrossRef]
  46. Shea, M.S. Hydraulic Conductivity of Cement-Treated Soils and Aggregates after Freezing. Master’s Thesis, Brigham Young University, Provo, UT, USA, 2011. [Google Scholar]
  47. Guthrie, W.S.; Shea, M.S.; Eggett, D.L. Hydraulic Conductivity of Cement-Treated Soils and Aggregates after Freezing. In Cold Regions Engineering 2012: Sustainable Infrastructure Development in a Changing Cold Environment; ASCE: Reston, VA, USA, 2012; pp. 93–103. [Google Scholar]
  48. Janoo, V.C.; Firicano, A.J.; Barna, L.A.; Orchino, S.A. Field Testing of Stabilized Soil. J. Cold Reg. Eng. 1999, 13, 37–53. [Google Scholar]
  49. Yarbaşı, N.; Kalkan, E.; Akbulut, S. Modification of the Geotechnical Properties, as Influenced by Freeze–Thaw, of Granular Soils with Waste Additives. Cold Reg. Sci. Technol. 2007, 48, 44–54. [Google Scholar] [CrossRef]
  50. Altun, S.; Goktepe, A.B. Effects of Freezing and Thawing Processes on the Mechanical Behavior of Silty Soils Stabilized with Fly Ash. In Advances in Transportation Geotechnics, 1st ed.; Ellis, E., Yu, H., McDowell, G., Dawson, A.R., Thom, N., Eds.; CRC Press: London, UK, 2008; pp. 636–641. [Google Scholar]
  51. Liu, J.; Wang, T.; Tian, Y. Experimental Study of the Dynamic Properties of Cement- and Lime-Modified Clay Soils Subjected to Freeze–Thaw Cycles. Cold Reg. Sci. Technol. 2010, 61, 29–33. [Google Scholar] [CrossRef]
  52. Kamei, T.; Ahmed, A.; Shibi, T. Effect of Freeze–Thaw Cycles on Durability and Strength of Very Soft Clay Soil Stabilised with Recycled Bassanite. Cold Reg. Sci. Technol. 2012, 82, 124–129. [Google Scholar]
  53. Shibi, T.; Kamei, T. Effect of Freeze–Thaw Cycles on the Strength and Physical Properties of Cement-Stabilised Soil Containing Recycled Bassanite and Coal Ash. Cold Reg. Sci. Technol. 2014, 106–107, 36–45. [Google Scholar] [CrossRef]
  54. Eskişar, T.; Altun, S.; Kalıpcılar, İ. Assessment of Strength Development and Freeze-Thaw Performance of Cement Treated Clays at Different Water Contents. Cold Reg. Sci. Technol. 2015, 111, 50–59. [Google Scholar] [CrossRef]
  55. Jamshidi, R.J.; Lake, C.B. Hydraulic and Strength Properties of Unexposed and Freeze–Thaw Exposed Cement-Stabilized Soils. Can. Geotech. J. 2015, 52, 283–294. [Google Scholar]
  56. Wang, S.; Lv, Q.; Baaj, H.; Li, X.; Zhao, Y. Volume Change Behaviour and Microstructure of Stabilized Loess under Cyclic Freeze–Thaw Conditions. Can. J. Civ. Eng. 2016, 43, 865–874. [Google Scholar] [CrossRef] [Green Version]
  57. Consoli, N.C.; da Silva, K.; Filho, S.; Rivoire, A.B. Compacted Clay-Industrial Wastes Blends: Long Term Performance under Extreme Freeze-Thaw and Wet-Dry Conditions. Appl. Clay Sci. 2017, 146, 404–410. [Google Scholar] [CrossRef]
  58. Lu, Y.; Liu, S.; Zhang, Y.; Li, Z.; Xu, L. Freeze-Thaw Performance of a Cement-Treated Expansive Soil. Cold Reg. Sci. Technol. 2020, 170, 102926. [Google Scholar] [CrossRef]
  59. Liu, C.; Berard, C.; Deng, L. Laboratory Investigation of Mechanical Behaviour of Cement-Treated Edmonton Clay Subjected to Freeze/Thaw Cycles. In Proceedings of the GeoVirtual—73rd conference of Canadian Geotechnical Society, Calgary, AB, Canada, 14–16 September 2020. [Google Scholar]
  60. Andersland, O.B.; Ladanyi, B. Frozen Ground Engineering; John Wiley & Sons. Inc.: Hoboken, NJ, USA, 2003. [Google Scholar]
  61. Wang, D.; Ma, W.; Niu, Y.; Chang, X.; Wen, Z. Effects of Cyclic Freezing and Thawing on Mechanical Properties of Qinghai–Tibet Clay. Cold Reg. Sci. Technol. 2007, 48, 34–43. [Google Scholar] [CrossRef]
  62. Tikalsky, P.J.; Pospisil, J.; MacDonald, W. A Method for Assessment of the Freeze–Thaw Resistance of Preformed Foam Cellular Concrete. Cem. Concr. Res. 2004, 34, 889–893. [Google Scholar] [CrossRef]
  63. Ababneh, A.N.; Xi, Y. Evaluation of Environmental Degradation of Concrete in Cold Regions. In Proceedings of the 13th International Conference on Cold Regions Engineering, American Society of Civil Engineers, Orono, ME, USA, 23–26 July 2006. [Google Scholar]
  64. Hale, W.M.; Freyne, S.F.; Russell, B.W. Examining the Frost Resistance of High Performance Concrete. Constr. Build. Mater. 2009, 23, 878–888. [Google Scholar] [CrossRef]
  65. Qi, J.; Ma, W.; Song, C. Influence of Freeze–Thaw on Engineering Properties of a Silty Soil. Cold Reg. Sci. Technol. 2008, 53, 397–404. [Google Scholar] [CrossRef]
  66. Xu, J.; Li, Y.; Lan, W.; Wang, S. Shear Strength and Damage Mechanism of Saline Intact Loess after Freeze-Thaw Cycling. Cold Reg. Sci. Technol. 2019, 164, 102779. [Google Scholar]
  67. ASTM. Test Method for Resistance of Concrete to Rapid Freezing and Thawing; American Society for Testing and Materials: West Conshohocken, PA, USA, 2016. [Google Scholar]
  68. ASTM. Standard Test Methods for Freezing and Thawing Compacted Soil-Cement Mixtures; American Society for Testing and Materials: West Conshohocken, PA, USA, 2016. [Google Scholar]
  69. Jamshidi, R.J.; Lake, C.B.; Gunning, P.; Hills, C.D. Effect of Freeze/Thaw Cycles on the Performance and Microstructure of Cement-Treated Soils. J. Mater. Eng. 2016, 28, 4016162. [Google Scholar]
  70. Wen, Y.Q.; Cui, C. Experimental Study on the Additive Effects of Cyclic Freezing-Thawing on the Durability of Cement Soil. Appl. Mech. Mater. 2014, 507, 363–367. [Google Scholar]
  71. ASTM. Standard Test Methods for Wetting and Drying Compacted Soil-Cement Mixtures; American Society for Testing and Materials: West Conshohocken, PA, USA, 2003. [Google Scholar]
  72. Jamshidi, R.J.; Lake, C.B.; Barnes, C.L. Physical Performance of Cement-Treated Silty Sand Soil under Cycles of Freezing Thawing. In Proceedings of the 14th Pan-American Conference on Soil Mechanics and Geotechnical Engineering (PCSMGE), the 64th Canadian Geotechnical Conference (CGC), Toronto, ON, Canada, 2–6 October 2011. [Google Scholar]
  73. Jamshidi, R.J.; Lake, C.B.; Barnes, C.L. Examining Freeze Thaw Cycling and Its Impact on the Hydraulic Performance of Cement-Treated Silty Sand. J. Cold Reg. Eng. 2015, 29, 4014014. [Google Scholar]
  74. Yousif, M.M.A. Freeze-Thaw Performance of Low-Cement Content Stabilized Soils for Containment Applications. Master’s Thesis, Dalhousie University, Halifax, NS, Canada, 2015. [Google Scholar]
  75. Lake, C.B.; Yousif, M.A.-M.; Jamshidi, R.J. Examining Freeze/Thaw Effects on Performance and Morphology of a Lightly Cemented Soil. Cold Reg. Sci. Technol. 2017, 134, 33–44. [Google Scholar] [CrossRef]
  76. Sezer, A.; Tanrinian, N.; Adamcil, Y.E. Mechanical Behavior of Fly Ash or Cement Amended Sand-Bentonite Mixtures Exposed to Freeze-Thaw Action after Long Curing. Anadolu Univ. J. Sci. Technol. A Appl. Sci. Eng. 2017, 18, 1008–1017. [Google Scholar] [CrossRef]
  77. Wang, Q.; Cui, J. Study on Strength Characteristics of Solidified Contaminated Soil under Freeze-Thaw Cycle Conditions. Adv. Civ. Eng. 2018, 2018, 8654368. [Google Scholar] [CrossRef] [Green Version]
  78. Shahin, M.A.; Jaksa, M.B.; Maier, H.R. Artificial neural network applications in geotechnical engineering. Aust. Geomech. 2001, 36, 49–62. [Google Scholar]
  79. Gullu, H.; Fedakar, H.I. On the Prediction of Unconfined Compressive Strength of Silty Soil Stabilized with Bottom Ash, Jute and Steel Fibers via Artificial Intelligence. Geomech. Eng. 2017, 12, 441–464. [Google Scholar] [CrossRef]
  80. Antemir, A.; Hills, C.D.; Carey, P.J.; Gardner, K.H.; Bates, E.R.; Crumbie, A.K. Long-Term Performance of Aged Waste Forms Treated by Stabilization/Solidification. J. Hazard. Mater. 2010, 181, 65–73. [Google Scholar] [CrossRef]
  81. Abbey, S.J.; Eyo, E.U.; Ng’ambi, S. Swell and Microstructural Characteristics of High-Plasticity Clay Blended with Cement. Bull. Eng. Geol. Environ. 2019, 79, 2119–2130. [Google Scholar] [CrossRef] [Green Version]
  82. Horpibulsuk, S.; Rachan, R.; Chinkulkijniwat, A.; Raksachon, Y.; Suddeepong, A. Analysis of Strength Development in Cement-Stabilized Silty Clay from Microstructural Considerations. Constr. Build. Mater. 2010, 24, 2011–2021. [Google Scholar] [CrossRef]
  83. Kolovos, K.G.; Asteris, P.G.; Cotsovos, D.M.; Badogiannis, E.; Tsivilis, S. Mechanical Properties of Soilcrete Mixtures Modified with Metakaolin. Constr. Build. Mater. 2013, 47, 1026–1036. [Google Scholar] [CrossRef]
  84. Güllü, H.; Agha, A.A. The Rheological, Fresh and Strength Effects of Cold-Bonded Geopolymer Made with Metakaolin and Slag for Grouting. Constr. Build. Mater. 2021, 12, 441–464. [Google Scholar]
  85. Güllü, H.; Al Nuaimi, M.M.D.; Aytek, A. Rheological and Strength Performances of Cold-Bonded Geopolymer Made from Limestone Dust and Bottom Ash for Grouting and Deep Mixing. Bull. Eng. Geol. Environ. 2021, 80, 1103–1123. [Google Scholar] [CrossRef]
  86. Bruce, M.E.C.; Berg, R.R.; Collin, J.G.; Filz, G.M.; Terashi, M.; Yang, D.S. Federal Highway Administration Design Manual: Deep Mixing for Embankment and Foundation Support (No. FHWA-HRT-13-046); Offices of Research & Development, Federal Highway Administration: Washington, DC, USA, 2013; p. 248.
  87. Ekinci, A.; Ince, C.; Ferreira, P.M.V. An Experimental Study on Compression and Shrinkage Behavior of Cement-Treated Marine Deposited Clays. Int. J. Geosynth. Ground Eng. 2019, 5, 1–17. [Google Scholar] [CrossRef]
  88. Perera, A.S.R.; Al-Tabbaa, A.; Reid, J.M.; Johnson, D. State of Practice Report UK Stabilisation Solidification Treatment and Remediation, Part V Long Term Performance and Environmental Impact. In Proceedings of the International Conference on Stabilization/Solidification Treatment and Remediation, Cambridge, UK, 12–13 April 2005; pp. 437–457. [Google Scholar]
  89. Paria, S.; Yuet, P.K. Solidification–Stabilization of Organic and Inorganic Contaminants Using Portland Cement: A Literature Review. Environ. Rev. 2006, 14, 217–255. [Google Scholar]
  90. De Silva, M.S.; O’Riordan, N.J.; Parry, L.N. Trials for the Construction of a Cement Solidified Retaining Structure in a Domestic Landfill Site Using Deep Soil Mixing. Eng. Geol. 2001, 60, 49–60. [Google Scholar] [CrossRef]
  91. Mengue, E.; Mroueh, H.; Lancelot, L.; Eko, R.M. Mechanical Improvement of a Fine-Grained Lateritic Soil Treated with Cement for Use in Road Construction. J. Mater. Eng. 2017, 29, 4017206. [Google Scholar] [CrossRef]
  92. Zheng, G.; Jiang, Y.; Han, J.; Liu, Y.-F. Performance of Cement-Fly Ash-Gravel Pile-Supported High-Speed Railway Embankments over Soft Marine Clay. Mar. Georesour. Geotechnol. 2011, 29, 145–161. [Google Scholar] [CrossRef]
  93. Timoney, M.J.; McCabe, B.A.; Bell, A.L. Experiences of Dry Soil Mixing in Highly Organic Soils. Proc. Inst. Civ. Eng. Ground Improv. 2012, 165, 3–14. [Google Scholar] [CrossRef]
  94. Kitazume, M.; Terashi, M. The Deep Mixing Method; CRC Press: London, UK, 2013. [Google Scholar]
  95. Tang, C.-S.; Pei, X.-J.; Wang, D.-Y.; Shi, B.; Li, J. Tensile Strength of Compacted Clayey Soil. J. Geotech. Geoenviron. Eng. 2015, 141, 04014122. [Google Scholar]
  96. Li, Y.; Guan, F.; Su, H.; Aydin, A.; Beroya-Eitner, M.A.; Zachert, H. A New Direct Tension Test Method for Soils and Soft Rocks. Geotech. Test. J. 2020, 43, 20190308. [Google Scholar] [CrossRef]
  97. Cao, B.; Souza, L.; Xu, J.; Mao, W.; Wang, F.; Al-Tabbaa, A. Soil Mix Cutoff Wall Materials with Microcapsule-Based Self-Healing Grout. J. Geotech. Geoenviron. Eng. 2021, 147, 04021124. [Google Scholar] [CrossRef]
  98. Cao, B.; Zhang, Y.; Xu, J.; Al-Tabbaa, A. Use of Superabsorbent Polymer in Soil-Cement Subsurface Barriers for Enhanced Heavy Metal Sorption and Self-Healing. Sci. Total Environ. 2022, 831, 154708. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Applications of soil mixing in cold regions: (a) oil tanks using DSM as the foundation, (b) retaining walls, (c) support for land reclamation, (d) soil mixing base to mitigate the potential seismic activity, (e) dam core to eliminate seepage, and (f) contaminant containment with the S/S method (drawings not to scale).
Figure 1. Applications of soil mixing in cold regions: (a) oil tanks using DSM as the foundation, (b) retaining walls, (c) support for land reclamation, (d) soil mixing base to mitigate the potential seismic activity, (e) dam core to eliminate seepage, and (f) contaminant containment with the S/S method (drawings not to scale).
Geotechnics 02 00047 g001
Figure 2. F/T setups: (a) a commonly used refrigerator and (b) a 3D F/T system (adapted from [70]).
Figure 2. F/T setups: (a) a commonly used refrigerator and (b) a 3D F/T system (adapted from [70]).
Geotechnics 02 00047 g002
Figure 3. Mass loss caused by F/T cycles: (a) soilcrete with different water content under 12 F/T cycles (adapted from [55]), (b) soilcrete with different cement content under 12 F/T cycles (adapted from [34]), and (c) mass loss increases with the increase in F/T cycles (adapted from [57]).
Figure 3. Mass loss caused by F/T cycles: (a) soilcrete with different water content under 12 F/T cycles (adapted from [55]), (b) soilcrete with different cement content under 12 F/T cycles (adapted from [34]), and (c) mass loss increases with the increase in F/T cycles (adapted from [57]).
Geotechnics 02 00047 g003
Figure 4. Permeability change caused by F/T cycles: (a) permeability versus water content for soilcrete after 12 F/T cycles (adapted from [73]) and (b) permeability versus water content for soilcrete after 3 F/T cycles (adapted from [75]).
Figure 4. Permeability change caused by F/T cycles: (a) permeability versus water content for soilcrete after 12 F/T cycles (adapted from [73]) and (b) permeability versus water content for soilcrete after 3 F/T cycles (adapted from [75]).
Geotechnics 02 00047 g004
Figure 5. UCS ratio versus F/T cycles (adapted from [8,52,58]).
Figure 5. UCS ratio versus F/T cycles (adapted from [8,52,58]).
Geotechnics 02 00047 g005
Figure 6. The effect of binders on the performance of soilcrete subjected to F/T cycles: (a) UCS of soilcrete subjected to F/T cycles with varying bassanite content [52], (b) UCS of soilcrete subjected to F/T cycles with varying fly ash content [50], (c) UCS of soilcrete subjected to F/T cycles with varying bentonite and fly ash content [76].
Figure 6. The effect of binders on the performance of soilcrete subjected to F/T cycles: (a) UCS of soilcrete subjected to F/T cycles with varying bassanite content [52], (b) UCS of soilcrete subjected to F/T cycles with varying fly ash content [50], (c) UCS of soilcrete subjected to F/T cycles with varying bentonite and fly ash content [76].
Geotechnics 02 00047 g006
Figure 7. DSM technique: (a) DSM blocks, (b) DSM columns, and (c) DSM panels (courtesy of Keller).
Figure 7. DSM technique: (a) DSM blocks, (b) DSM columns, and (c) DSM panels (courtesy of Keller).
Geotechnics 02 00047 g007
Figure 8. Construction of DSM columns with the wet mixing method: (a) before drilling and (b) retrieval and mixing, and (c) block construction by wet mixing.
Figure 8. Construction of DSM columns with the wet mixing method: (a) before drilling and (b) retrieval and mixing, and (c) block construction by wet mixing.
Geotechnics 02 00047 g008
Figure 9. Specimen preparation: (a) lumped stiff clay in a container, (b) making cement slurry, (c) mixing slurry with soils, (d) lubricating the plastic molds, (e) curing in a rubber container filled with water, (f,g) cylinder specimens with a diameter of 75 mm, and (h,i) cylinder specimens with a diameter of 50 mm.
Figure 9. Specimen preparation: (a) lumped stiff clay in a container, (b) making cement slurry, (c) mixing slurry with soils, (d) lubricating the plastic molds, (e) curing in a rubber container filled with water, (f,g) cylinder specimens with a diameter of 75 mm, and (h,i) cylinder specimens with a diameter of 50 mm.
Geotechnics 02 00047 g009
Figure 10. Temperature control system: (a) steel container where the samples were kept, (b) samples kept in the cell, (c) the cell containing samples and sand wrapped for insulation, and (d) chiller used to conduct the heat transfer.
Figure 10. Temperature control system: (a) steel container where the samples were kept, (b) samples kept in the cell, (c) the cell containing samples and sand wrapped for insulation, and (d) chiller used to conduct the heat transfer.
Geotechnics 02 00047 g010
Figure 11. Photos of soilcrete samples before and after F/T cycles: (a) C2-D56T5N1, (b) C2-D56T2N1, (c) C2-D56T10N1, and (d) C2-D56T5N4.
Figure 11. Photos of soilcrete samples before and after F/T cycles: (a) C2-D56T5N1, (b) C2-D56T2N1, (c) C2-D56T10N1, and (d) C2-D56T5N4.
Geotechnics 02 00047 g011
Figure 12. Void ratio calculated from CT scans: (a,b) cross-section slices from different locations and (c) void ratio distribution along the sample (adapted from [9]).
Figure 12. Void ratio calculated from CT scans: (a,b) cross-section slices from different locations and (c) void ratio distribution along the sample (adapted from [9]).
Geotechnics 02 00047 g012
Table 2. Key references on the performance of soilcrete subject to F/T cycles.
Table 2. Key references on the performance of soilcrete subject to F/T cycles.
ReferenceApplicationSoil TypeBindersBinder DosageTest MethodsF/T MethodF/T CyclesComments
[48]SubgradeCrushed concrete and natural soilCement3.5%dynamic
cone penetrometer, Clegg hammer
Field conditionN/A50% of compressive strength loss after F/T cycles
[34]PavementGranular soilsCement5, 7, 9%UCS, brushing testFreezer, 3D, 22 h + 22 h12Proposed relationship between mass loss and decrease in UCS
[49]Road construction, earthquake applicationSand–gravel mixtureSilica fume, fly ash, lime, red mud, cement2.5 to 20%UCS, CBR, ultrasonic wave, dynamic testsFreezer, 3D, ASTM C 666, 2.3 h + 2.3 h60Addition of waste materials reduced the compressive strength loss significantly
[50]PavementMLCement, fly ash2 to 30%UCS Freezing cabinet, 3D 24 h + 23 h, ASTM D5607Addition of 10% fly ash and 2% cement has the best performance against F/T cycles
[51]Base layers of high-speed railwayNortheast China clay soilCement and calcium lime3 to 12%Cyclic loadingFreezer, 3D, 12 h + 12 h15Cement improved the durability more than lime with the same content
[52]Highway embankmentCHBassanite, furnace cement5 to 30%UCS, volume, WCFreezer, 3D, 24 h + 24 h5Early F/T cycles caused more damage than later ones
[53]EarthworksCHBassanite, furnace cement, coal ash5 to 20%UCS, volume, WCFreezer, 3D, 24 h + 24 h5Coal ash only does not result in improvement
[54]StabilizationCH and CLCement 5 and 10%UCS, ultrasonic waveFreezer, 3D, ASTM C 666, 2.3 h + 2.3 h5UCS prediction equation using clay water to cement ratio; the relationship between UCS and ultrasonic wave velocity
[55]StabilizationArtificial soils: well-graded sand and silty sandCement 10%Permeability, UCS, brushing testFreezer, 3D, 24 h + 24 h12Permeability increased significantly; compressive strength decreased by over 50%
[56]SubgradeLishi loessCement,
lime, and fly ash
3 to 10%Frost heave and thaw shrinkage, SEMFreezer, 3D, 12 h + 12 h8The contents resulting in the best improvement are 5% cement, 6% lime, and 10% fly ash
[57]EarthworkCHCoal fly ash, carbide lime3 to 9%UCS, mass loss, XRDFreezing cabinet, 3D 24 h + 23 h, ASTM D56012Mass loss and porosity/lime index to indicate the UCS
[39]High-speed railway subgradeMacadamCement 1, 3, and 5%Compaction, permeability, UCSA special device, 1D, 48 h + 48 h25UCS becomes stable after 10 F/T cycles; grain-size composition affects permeability, UCS, and frost sensitivity
[58]SNWTPexpansive clayCement 3, 5, 7%UCS, volume change, mass lossFreezer, 3D, 12 h + 12 h12Cement reduces the volume change of expansive clay
[16]StabilizationSilty sandCement and cement kiln dust6 and 8%UCSFreezer, 3D, 12 h + 36 h124 °C curing; healing effect
Note: CBR—California bearing ratio; SNWTP—south to north water transfer project; ML—silt.
Table 3. Key references on the microscale structure of soilcrete.
Table 3. Key references on the microscale structure of soilcrete.
ReferenceApplicationSoil TypeBindersBinder DosageTest MethodsF/T MethodF/T CyclesComments
[10]Marine clayStabilizationCement10%, 30%, 50%XRD, SEM, MIP, PSD by laser diffractometricNANAMicrostructure mechanisms: flocculation of the illite clay particles, surface deposition and shallow infilling, etc.
[53]Earthwork projectsclay with high plasticityCement, bassanite, coal ash5 to 20%UCS, SEMFreezer, 3-D, 24 h + 24 h5Quantity and size of ettringite are improved by mixing bassanite and coal ash because of the pozzolanic reaction
[74]StabilizationSilty sandCement 3 and 6%UCS, permeability, MIP, optical microscope, the resonant frequencyFreezer, 3-D, 24 h + 24 h3Damage in permeability is due to cracking and macroscale pore changes
[69]StabilizationSilty sandCement 10%Permeability, UCS, optical microscope, longitudinal impact resonance testFreezer, 1-D and 3-D, 24 h + 24 h3Optical microscopy can detect matrix disintegration for highly damaged specimens
[75]S/SSilty sandCement3 and 6%UCS, optical microscope, MIPFreezer, 3-D, 24 h + 24 h3 Thin section observations showed more voids and cracks in F/T exposed specimens relative to control samples
Table 4. Natural soil characteristics.
Table 4. Natural soil characteristics.
PropertyValue
Sand content33%
Silt content43%
Clay content24%
Unified Soil Classification System (USCS)Sandy CL
Water content21.8%
Liquid limit40.9%
Plastic limit12.5%
Plasticity index28.4%
Shear strength63–72 kPa (vane shear), 50 kPa (UCS)
Specific gravity2.54
Table 5. Parameters of different specimens.
Table 5. Parameters of different specimens.
PropertyC1C2 and CS2C3
Wc [%]55.5753.8452.14
Cc [%]17.5022.3927.47
Wc/Cc3.182.401.90
Binder content [kg/m3]175225275
Note: Wc: water mass to solids mass; Cc: cement mass to solids mass; binder content: binder mass to the volume of natural soils and slurry.
Table 6. Test matrix.
Table 6. Test matrix.
TypeCuring (D, Day)Temperature (T, °C)Number of F/T Cycles (N)Test Methods
C225 (C2 and CS2)1, 7, 14, 28, 110, 300−2, −5, −100, 1UCS, CU, Permeability, Brazilian test (BT), CT, SEM
56−2, −100, 1
−50 to 20
C175 (C1) and C275 (C3)14, 28, 56, 110, 300−2, −5, −100, 1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, C.; Berard, C.; Deng, L. Review of the Effects of Freezing/Thawing Cycles on the Engineering Properties of Soilcrete. Geotechnics 2022, 2, 1003-1026. https://0-doi-org.brum.beds.ac.uk/10.3390/geotechnics2040047

AMA Style

Liu C, Berard C, Deng L. Review of the Effects of Freezing/Thawing Cycles on the Engineering Properties of Soilcrete. Geotechnics. 2022; 2(4):1003-1026. https://0-doi-org.brum.beds.ac.uk/10.3390/geotechnics2040047

Chicago/Turabian Style

Liu, Chao, Claude Berard, and Lijun Deng. 2022. "Review of the Effects of Freezing/Thawing Cycles on the Engineering Properties of Soilcrete" Geotechnics 2, no. 4: 1003-1026. https://0-doi-org.brum.beds.ac.uk/10.3390/geotechnics2040047

Article Metrics

Back to TopTop